首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Translational diffusion, velocity sedimentation and viscometry of polyamidobenzimidazole (PABI) solutions in the range of M = (1–61) · 103 have been investigated in N,N-dimethylacetamide (DMA) and 98% H2SO4. The dependences of D0, S0 and [η] on M were obtained. Tsvetkov-Klenin's hydrodynamic invariant was found to be A0 = 3.55 · 10?10erg deg?1mol?13. The equilibrium rigidity of PABI molecules was characterized by the length of the Kuhn segment A = 250 ± 100 A?. The chain diameter was 7 ± 4 A?. The values of A in 98% H2SO4 and in an aprotic solvent, DMA, were virtually identical, implying that the rigid-chain conformation of PABI molecules in 98% H2SO4 is due to their geometrical structure rather than to the protonization of amide bonds. The significance of the latter evidently increases in PABI solutions in 100% H2SO4 in which A is 1.5 times as high. The decrease in rigidity of PABI as compared to that of poly-p-phenylene terephthalamide (A = 400 ± 100 A?) is in reasonable agreement with the presence of imidazole rings in PABI molecules. The presence of these rings results in kinks in the PABI chain with angles of about 30° and hence, in the depature from parallelism of rotating bonds.  相似文献   

2.
Flow birefringence (FB) and intrinsic viscosity of 19 samples of aromatic polyamide hydrazide (PAH) in dimethylsulphoxide (DMS), previously characterized by their weight-average molecular weights by the light scattering method, have been investigated. The molecular-weight dependence of reduced birefringence according to theory [12] was used to determine the optical anisotropy of a monomer unit Δa = (200 ± 20) 10?25cm3 and the length of the Kuhn segment A = (250 ± 30) A? of PAH molecules. The second independent evaluation of rigidity of the PAH chain A = (240 ± 30) A? was obtained according to the theory of rotational friction of rigid wormlike chains by using the coefficients of rotational diffusion of PAH molecules determined from the characteristic values of orientation angles of FB. The value of rigidity of the PAH chain obtained by this method is in good agreement with the data on molecular dimensions obtained by light scattering.  相似文献   

3.
Translational diffusion, velocity sedimentation and viscosity in acetone as well as flow birefringence (FB) and viscosity in cyclohexanone have been investigated for cyanoethyl cellulose (CEC) with degree of substitution 2.6 in the range of M = (24.5?317) × 103. The dependences of [ν], So and Do on M were obtained. The value of the hydrodynamic constant is A0 = 3.27 × 10?10erg deg?1mol?13. According to hydrodynamic data, the equilibrium rigidity of CEC molecules is characterized by the length of the Kuhn segment A = 240 ? 350 A? and the coefficient of hindrance to intramolecular motion σ = 4.5-5.4. The hydrodynamic diameter of the chain is 8–14 Å. According to the FB data, the value of A is 260 Å. This value is in agreement with hydrodynamic data. The high value of optical anisotropy of the monomer unit, a| - a = 17.8 × 10?25 cm3, is in agreement with the structure and anisotropy of the substituting groups, and the investigation of orientation angles of FB leads to the conclusion that, apart from high equilibrium rigidity, CEC in solution is characterized by considerable kinetic chain flexibility. The data for CEC are compared with the characteristics of other cellulose esters and ethers.  相似文献   

4.
The theta temperature for the system poly(o-chlorostyrene)-methyl ethyl ketone has been determined as 24·5°. The samples used in the determination were prepared by radical polymerization. The dependence of intrinsic viscosity on molecular weight has been measured in methyl ethyl ketone at 24·5° and found to be ηθ = 4·68 × 10?4MwM12. The ratio 〈s=2〉/M was found, by light scattering, to be 5·60 × 10?18 cm2. Analysis of the solution properties indicates that the Kurata-Yamakawa theory is valid in the vicinity of the Flory temperature (UCST).  相似文献   

5.
The heat capacity of the solid solution Mn3.2Ga0.8N was measured between 5 to 330 K by adiabatic calorimetry. A sharp anomaly with first-order character was detected at TA = (160.5±0.5) K, corresponding to a magnetic rearrangement and a lattice expansion. No sharp anomaly was observed at Tc ≈ 260 K where the magnetic ordering takes place; instead, a smooth shoulder was detected. The thermodynamic functions at 298.15 K are Cp,mR = 15.16, SmoR = 18.57, {Hmo(T)?Hmo(0)}R = 2896 K, ?{Gmo(T)?Hmo(0)}RT = 8.85. At low temperatures the coefficient for the linear electronic contribution to the heat capacity was derived: γ = (0.031±0.003) J·K?2·mol?1. Moreover, the different contributions to the heat capacity were obtained and the electronic origin of the phase transitions was established.  相似文献   

6.
The structure of GdPS is orthorhombic, space group Pmnb; a = 5.3620(5), b = 5.4079(6), c = 16.742(2)Å, Z = 8. It is a distorted derivative of the tetragonal PbFCl structure (a0, c0) with a ≈ b ≈ 212a0, c = 2c0. The distortions are due to the formation of phosphorus chains. This structure type is found also in other rare-earth sulfopolyphosphides, e.g., with Ln = La···Sm, Tb···Tm, Y.  相似文献   

7.
The tracer diffusion coefficient, D1O, of oxide ions in LaCoO3 single crystal was determined over the temperature range of 700–1000°C by a gas-solid isotopic exchange technique using 18O tracer. For the determination, two methods, the gas phase analysis and the depth profile measurement, were employed. Under an oxygen pressure of 34 Torr, the temperature dependence of D1O in LaCoO3 was expressed by
D1O(cm2·sec?1) = 3.63 × 104exp? (74 ± 5)kcal · mole?1RT
D1O at 950°C was found to be proportional to P?0.35O2. The diffusion of oxide ions occurs through a vacancy mechanism. The activation energy for the migration of oxide ion vacancies was estimated as 18 kcal · mole?1.  相似文献   

8.
For a number of fractions and unfractionated samples of polylaurolactam, molecular weights (Mw = 1 × 104?12·5 × 104) were measured by the light-scattering method in a mixed solvent of m-cresol with 60 vol. % 2,2,3,3-tetrafluoropropanol; intrinsic viscosities were determined in m-cresol and 96% H2SO4, and the constants of the Mark-Houwink equation were calculated. The calculated values of the characteristic ratio of unperturbed dimensions (virtually identical for m-cresol and 96% H2SO4) were compared with the respective values for polypyrrolidone and polycaprolactam. It was found that the higher frequency of the CONH-groups reduces the rigidity of the polyamide chain.  相似文献   

9.
The compounds Ba4Fe2S6[S23(S2)13] and Ba3.6Al0.4Fe2S6[S0.6(S2)0.4], designated I and II, were prepared by reacting BaS, Fe, and S powders and Al foils in graphite containers sealed in evacuated quartz ampoules at approximately 1100°C. The crystal structure of I was determined using 1682 independent, nonzero X-ray reflections, while 3589 were used for II. They are triclinic, Al:
a=9.002(2)A?,b=6.7086(8)A?,c=24.658(4)A?α91.49(2)°,
β=105.10(2)°y=90.74(2)°,ψcalc=4.15g/cm3,for I:
a=8.993(6)A?,b=6.708(7)A?,c=24.70(1)A?α91.11(6)°,
β=105.04(6)°y=90.90(9)°,ψcalc=3.90g/cm3,for II:
BaS6 trigonal prisms share edges to form distorted hexagonal rings which form one-dimensional chains leaving two free lateral edges. The chains link in a stairstep manner with the rings offset along the [301] direction. These stairsteps join in a complicated manner to form a three-dimensional network. Fe ions are in two sites forming isolated FeS4 tetrahedra and isolated Fe2S6 dimers by edge-sharing tetrahedra. The Al substitution occurs in the trigonal prisms which have free edges with Al replacing Ba. Room-temperature Mössbauer isomer shifts are 0.20 mm/sec. for I and 0.30 mm/sec for II. These data indicate that upon Al substitution charge compensation occurs by reducing Fe3+. Valence calculations indicate that Fe in edge-sharing tetrahedra are reduced while the Fe in the isolated tetrahedron remains unchanged. The effective charge distribution in the Al substituted compound is approximately Fe3+, Fe2.5+ with electron delocalization across the shared edge. Room temperature electrical resistivity is 105 ohm/cm. The compositions of the crystals are best represented by the formulas [Ba4Fe2S7]23·[Ba4Fe2S6(S2)]13 and [Ba3AlFe2S7]0.4·[Ba4Fe2S7]0.2·[Ba4Fe2S6(S2)]0.4. The replacement of a sulfide by a disulfide ion is thought to be strongly dependent on the sulfur activity during the preparation.  相似文献   

10.
The electrical conductivity of polycrystalline strontium titanate with (SrTi = 0.996, 0.99, and 0.98 was determined for the oxygen partial pressure range of 100 to 10?22 atm and the temperature range of 850–1050°C. These data were found to be similar to that obtained for the sample with ideal cationic ratio. The observed data were proportional to the ?16 power of oxygen partial pressure for PO2 < 10?15atm, proportional to P?14O2 for the pressure range 10?8–10?15 atm, and proportional to P+14O2 for PO2 > 10?4atm. The deviation from the ideal Sr-to-Ti ratio was found to be accommodated by neutral vacancy pairs, (V″Sr V″0. The results indicate that the single-phase field of strontium titanate extends beyond 50.505 mole% TiO2 at elevated temperatures.  相似文献   

11.
The electrical conductivity of polycrystalline CaTiO3 was measured over the temperature range 800–1100°C while in thermodynamic equilibrium with oxygen partial pressures from 10?22 to 100 atm. The data were found to be proportional to the ?16th power of the oxygen partial pressure for the oxygen pressure range 10?16 – 10?22 atm, proportional to P?14O2 for the oxygen pressure range 10?8 – 10?15 atm, and proportional to P+14O2 for the oxygen pressure range greater than 10?4 atm. The region of linearity where the electrical conductivity varies as ?14th power of PO2 increased as the temperature was decreased. The observed data are consistent with the presence of small amounts of acceptor impurities in CaTiO3. The band-gap energy (extrapolated to zero temperature) was estimated to be 3.46 eV.  相似文献   

12.
The electrical conductivity of polycrystalline SrTiO3 was determined for the oxygen partial pressure range of 10° to 10?22 atm and temperature range of 800 to 1050°C. The data were found to be proportional to the ?16th power of the oxygen partial pressure for the oxygen pressure range 10?15–10?22 atm, proportional to P?14O2 for the oxygen pressure range 10?8–10?15 atm, and proportional to P+14O2 for the oxygen pressure range 100–10?3 atm. These data are consistent with the presence of very small amounts of acceptor impurities in SrTiO3.  相似文献   

13.
The kinetics, mechanism, and activation energy of the isothermal decomposition of CuCrO4 was studied using an isothermal TG method and an X-ray high-temperature diffraction technique in either air or a flowing atmosphere of N2. The enthalpy change ΔH of the decomposition reaction
2CuCrO4CuO+CuO+CuCr2O4+32O2
was determined by DSC analysis. The mechanism of the thermal decomposition of CuCrO4 is well represented by the standard Avrami-Erofeev kinetic equation [?ln(1 ? α)]12 = kt. According to this mechanism, the reaction rate is controlled by the formation and growth of nuclei on the surface of the reactant. The activation energy EA of the process in air is EA = (248 ± 8) kJ mole?1, in flowing atmosphere of nitrogen EA = (229 ± 8) kJ mole?1. ΔH in air is 110 kJ mole?1, in flowing nitrogen 67 kJ mole?1. The lower values of ΔH and EA in the flowing atmosphere of nitrogen are due to the fast elimination of O2 from the reaction interface. However, the decay of the crystalline portion of CuCrO4 during its thermal decomposition, studied by the X-ray diffraction, is controlled by a different reaction mechanism (first-order kinetics). The reaction mechanism is discussed in the relation to the crystal structure of the reactants.  相似文献   

14.
The electrical conductivity and departure from the stoichiometry of Nd2O3 have been measured over the temperature range of 900° to 1100°C and oxygen partial pressure of 1 to 10?16 atm. The hole conductivity of Nd2O3 is found to be proportional to P1nO2, where n are 4.6, 4.9, and 5.1 at 900°, 1000°, and 1100°C, respectively. From the oxygen partial pressure dependence of the hole conductivity, it is shown that the predominant point defects in nonstoichiometric NdO1·+x are fully ionized and partially doubly ionized metal vacancies. From the thermogravimetric measurements, the departure from stoichiometry, x in NdO1·5+x, is 2.0 × 10?3 at 1000°C and 1 atm. By combining the electrical conductivity and weight change data, it is shown that the hole mobility is 6.3 × 10?4 (cm2/V·sec) at 1000°C and 1 atm.  相似文献   

15.
The kinetics of the interaction of hexaaquochromium(III) ion with potassium octacyanomolybdate(IV) have been studied using conductance and spectrophotometric data. The mechanism of the reaction is discussed and the effect of H+ ion and the ionic strength on the rate of the reaction determined. The reaction is found to be pseudo-first order with respect to potassium octacyanomolybdate(IV) and inverse first order with [H3O+]. The rate of the reaction increases with increase in ionic strength and temperature. Activation parameters have been calculated using the Arrhenius equation and have the values ΔE* = 1.3 × 102 kJ mol?1, ΔH* = 129 kJ mol?1, ΔS* = ?315 e.u., ΔF* = 2.3 × 102 kJ and A = 1.5 × 10?3. The mechanism proposed is based on ion-pair formation and the rate equation obtained is given by: kobs=[kKE[H3O+]+k′K′kEkh][Mo(CN)84?][H3O+]+kh+[KE[H3O+]+K′Ekh][Mo(CN)84?]  相似文献   

16.
The polymerization of acrylamide, initiated with permanganate/oxalic acid, has been studied at 35 ± 0.2 in an aqueous medium under nitrogen. Samples of polyacrylamide were fractionated by the triangular fractionation method using methanol as non-solvent. Molecular weights of the fractions have been determined by viscometry and osmometry. Integral and differential distribution curves were plotted using the fractionation data. The narrow molecular weight distribution for high conversion polymers has been discussed. For fractionated samples of polyacrylamide in water at 30°, the equation [η]ml/g = 6.5 × 10?3Mn0.82 is applicable for the molecular weight range 4 × 104 to 127 × 104. This equation is very similar to the equation [η]ml/g = 6.31 × 10?3M/0.80 of Scholtan. Other parameters (osmotic second virial coefficient and unperturbed dimension) have also been evaluated.  相似文献   

17.
The polymerization of styrene initiated by 2,2′ azobisisobutyronitrile had been studied in N,N-dimethylformamide at 60°, in the presence of Tris(phenanthroline)-iron(III) perchlorate. The complex was prepared in situ by mixing phenanthroline with hexakis (N,N-dimethylformamide) iron(III) perchlorate in the ratio 3:1. The nature of the complex formed was established by Job's method. The equilibrium constant for
Fe3+ + 3 Phen ? [Fe(Phen)3]3+
is 2·3 × 102 13 mol?3. The velocity constant at 60° for the reaction of polystyryl radical with [Fe(Phen)3]3+ is 2·93 × 104 mol?1 l s?1.  相似文献   

18.
The study of K2NiF4 and perovskite structure type by the “method of invariants” leads to the relationship: (A-X)9 212 ? (A-X)12 = constant, where (A-X)9 and (A-X)12 are the invariant values associated with cation A in coordination number 9 and 12. In the case where A = K+ and X = F?, we propose the relationship:
(K+?F)R = 2.832 R111.4
where R is the coordination number.  相似文献   

19.
Photon-correlation spectroscopy has been used to measure the diffusion coefficient D and first-mode intramolecular relaxation time τ1 of polystyrene in a theta solvent, cyclohexane at 34.5°C. Measurements were made on five narrow fractions (Mw = 2.88 × 106 to 9.35 × 106) as a function of concentration c, in the dilute regime. D varied linearly with c, D = Do (I + kDc), with Do = (1.4 ± 0.2) × 10?4Mw?(0.508±0.007) cm2 s?1. Although the values obtained for τ1 were reproducible to within 5%, no systematic variation with c was detected. The results are fitted by the relation τ1 = (7.7 ? 0.3) × 10?8Mw(1.42±0.05) μs, which agrees well with the theoretical expression of Zimm for the non-draining bead-and-spring model, modified for the light-scattering case.  相似文献   

20.
The kinetics of the anionic polymerization of styrene initiated by cumylcaesium have been reinvestigated in tetrahydrofuran. Quite unlike earlier investigations, the ion pair rate constant kp(±) was found to be 130 l/mole/sec at 25° and the dissociation constant of polystyrylcaesium to be about 5 × 10?10 mole/l. In the presence of caesiumtriphenylcyanoborate used as a common ion salt addition, kp(±) was measured over a wide range of temperature. The results are compared to the temperature behaviour of the ion pair rate constants obtained with polystyrylsodium in polar solvents; they indicate that the existence of solvent separated ion pairs cannot be excluded for polystyrylcaesium.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号